β-FeOOH: An Earth-Abundant High-Capacity Negative Electrode

Jul 20, 2015 - ABSTRACT: Thanks to the great earth abundance and excellent energy density of sodium, sodium-ion batteries are promising alternative ...
0 downloads 0 Views 2MB Size
Page 1 of 11

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

β-­‐FeOOH:  An  Earth-­‐Abundant  High-­‐Capacity  Negative  Electrode   Material  for  Sodium-­‐Ion  Batteries   Linghui  Yu,†  Luyuan  Paul  Wang,†,  ‡  Shibo  Xi,§  Ping  Yang,⊥    Yonghua  Du,§,*  Madhavi  Sriniva-­‐ san,†,§  and  Zhichuan  J.  Xu†,§,*   †

 School  of  Materials  Science  and  Engineering,  Nanyang  Technological  University,  50  Nanyang  Avenue,  639798,   Singapore   §

 Energy  Research  Institute@NTU,  Nanyang  Technological  University,  50  Nanyang  Drive,  639798,  Singapore  



 Institute  of  Chemical  and  Engineering  Sciences,  A*STAR,  1  Pesek  Road,  627833,  Singapore  

⊥  Singapore  Synchrotron  Light  Source  (SSLS),  National  University  of  Singapore,  5  Research  Link,  117603,  Singapore.  

  ABSTRACT:  Thanks  to  the  great  earth  abundance  of  sodium  and  decent  energy  densities,  sodium-­‐ion  batteries  are  prom-­‐ ising  alternative  energy  storage  devices  for  large-­‐scale  applications.  Developing  cheap,  safe  and  high-­‐capacity  sodium-­‐ion   battery  anode  materials  is  one  of  critical  challenges  in  this  field.  Here,  we  show  that  β-­‐FeOOH  is  a  very  promising  low-­‐ cost  anode  material,  with  a  high  reversible  capacity  (>500  mAh  g-­‐1  during  initial  cycles).  The  fundamental  characteristics   associated  with  the  discharge/charge  processes,  in  terms  of  the  redox  reactions,  formation/deformation  of  the  solid  elec-­‐ trolyte  interface  (SEI)  layers,  and  structural  and  morphological  changes,  are  comprehensively  investigated.  In  addition,  a   comparison  study  shows  that  the  smaller-­‐sized  FeOOH  has  more  serious  kinetic  restrictions,  and  thus  lower  capacities,   while  it  shows  better  cyclability  than  the  bigger  one.  Origins  of  the  large  overpotential  are  discussed  and  it  is  suggested   that  the  overpotential  should  be  mainly  due  to  the  features  of  the  surface  concentration  dependent  potential  and  the  slow   diffusion  of  Na+;  and  the  presence  of  the  SEI  layers  may  also  contribute  to  the  overpotential.    

INTRODUCTION   Sodium   (Na)-­‐ion   batteries   (NIBs)   have   been   receiving   increasing   research   attention   in   recent   years,   due   to   the   potential  use  for  large-­‐scale  applications,  such  as  station-­‐ ary  energy  storage.1-­‐5  Compared  with  lithium-­‐ion  batteries   (LIBs),   NIBs   may   have   lower   but   still   considerably   high   energy   densities.   However,   the   great   abundance   of   sodi-­‐ um   makes   NIBs   potentially   much   cheaper,   hence,   very   attractive   for   large-­‐scale   applications   where   both   cost   and   energy  density  are  key  factors.   The   development   of   NIBs   has   been   snagged   perhaps   mainly   by   the   lack   of   suitable   electrode   materials.   Alt-­‐ hough   NIBs   are   conceptually   identical   to   LIBs,   the   large   radius  of  Na+  (0.3  Å  larger  than  that  of  Li+)  is  an  intrinsic   drawback  that  results  in  sluggish  diffusion  kinetics  of  Na+   in  electrode  materials,  leading  to  lower  capacities  or  even   electrochemical   inactivity.1,  6-­‐9   Intensive   efforts   have   been   devoted   to   improving   and   exploring   both   cathodes   and   anodes.   For   anodes,   hard   carbons   are   early   reported   ma-­‐ terials   with   significant   sodium   storage   capacities   at   low   potentials   vs.   Na+/Na.10-­‐14   So   far,   much   progress   has   been   made   in   terms   of   capacity   and   cycalibility.   Even   so,   they   seems   to   approach   a   limiting   value   around   300   mAh   g-­‐1.   Titanium-­‐based   materials,   such   as,   Na2Ti3O7,15-­‐18   Li4Ti5O12,19-­‐23   Na0.66[Li0.22Ti0.78]O2,24   TiO2,25-­‐28   and   Na-­‐ Ti2(PO4)3,29-­‐31  have  emerged  as  a  promising  class  of  anodes   for   NIBs.   Although   they   have   a   relatively   low   capacity  

between   100   –   300   mAh   g-­‐1,   several   of   them   have   shown   excellent  cylability  and  have  been  proposed  for  long-­‐term   usage.   Research  work  on  high-­‐capacity  anode  materials  focus-­‐ es  mainly  on  tin-­‐  and  antimony-­‐based  materials,  such  as,   Sn,32-­‐35   Sb,36-­‐40,   SnSb,41   SnO2,42,   43   SnS,7,   44   SnS2,7,   45,   46   Sb2O4,47   and   Sb2S3.48   With   the   electrochemical   formation   of   Na-­‐Sn   or   Na-­‐Sb   alloys,   they   own   high   theoretical   ca-­‐ pacities,   for   instance,   847   and   660   mAh   g-­‐1   for   pure   tin   and   antimony,   respectively.   Many   reports   have   shown   that  a  large  proportion  of  the  theoretical  capacities  could   be  realized,  from  500  mAh  g-­‐1  up  to  close  to  the  theoreti-­‐ cal   capacities.   Despite   the   high   capacity,   these   materials   usually  exhibit  poor  cyclability  and  there  are  concerns  on   the   costs   of   NIBs   raised   by   using   relatively   expensive   tin   or  antimony.  Their  earth  abundance  may  also  fail  to  meet   large-­‐scale   grid   applications.   Phosphorus-­‐based   materials,   such   as   P,49-­‐51   Sn4P3,52-­‐54   and   Ni3P,55   are   another   class   of   high-­‐capacity   anodes   reported   for   NIBs.   For   pure   phos-­‐ phorus,  it  has  the  highest  theoretical  capacity  (2596  mAh   g-­‐1)  of  all  the  anode  materials  reported.  And  a  capacity  up   to   1800   –   1900   mAh   g-­‐1   has   been   achieved   by   several   groups,  though  fades  fast  upon  cycling.  However,  the  use   of  phosphorus-­‐base  materials  involves  formation  of  highly   toxic   Na3P,56-­‐58   which   additionally   increases   safety   haz-­‐ ards.  

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

While   the   development   of   high-­‐capacity   alloying   and   phosphorus-­‐based   materials   is   accompanied   by   cost   and/or  safety  considerations,  the  use  of  iron-­‐based  mate-­‐ rials  becomes  highly  attractive,  because  of  their  low  costs,   environmental   friendliness,   and   high   theoretical   capaci-­‐ ties.   Iron-­‐based   materials,   such   as   Fe2O3   and   Fe3O4,   have   been   widely   investigated   for   LIBs.59,   60   They   can   electro-­‐ chemically   react   with   lithium   to   form   metallic   iron   and   deliver  high  theoretical  capacities,  e.g.,  1005  and  926  mAh   g-­‐1   for   Fe2O3   and   Fe3O4,   respectively.   Similar   mechanisms   of  formation  of  metallic  iron  have  been  also  proposed  for   their   sodium   storage.61-­‐63   But   the   sodium   storage   capaci-­‐ ties   are   usually   limited   by   the   slow   kinetics   associated   with  the  large  radius  of  Na+.  Among  the  reported  materi-­‐ als,  several  relatively  high  capacities  were  obtained  based   on   nano-­‐sized   particles   in   composites   with   carbons.   For   example,   a   Fe3O4   with   particle   size   less   than   10   nm   in   a   matrix   of   carbon   showed   a   capacity   of   ~350   mAh   g-­‐1   at   a   low   current   density.64   Fe2O3   nanocrystals   anchored   on   graphene  could  have  a  stable  capacity  around  400  mAh  g-­‐1   at   100   mA   g-­‐1.65   A   5   nm   γ-­‐Fe2O3   embedded   in   a   matrix   of   carbon   exhibited   a   high   capacity   that   could   be   stabilized   at   ~700   mAh   g-­‐1   at   100   mA   g-­‐1.66   So   far,   in   addition   to   ki-­‐ netic   problems,   iron-­‐based   materials   always   show   very   low  initial  reversibility.  Most  of  them  have  an  initial  Cou-­‐ lombic   efficiencies   (CEs)   of   lower   than   60%,62,   64,   65,   67,   68     while  some  of  them  have  a  value  close  to  70%.63  Such  low   CEs  are  unacceptable  for  practical  applications.   Herein,   we   present   a   promising   sodium   storage   anode   β-­‐FeOOH   with   a   high   reversible   capacity   (627   and   523   mAh   g-­‐1   for   the   first   and   second   cycle,   respectively,   at   a   current  of  80  mAh  g-­‐1).  The  sodium  storage  processes  are   investigated   by   a   combination   of   electrochemistry,   X-­‐ray   absorption  near  edge  structure  (XANES)  spectroscopy,  X-­‐ ray   diffraction   (XRD)   and   electron   microscopy   to   under-­‐ stand  the  redox  reactions,  formation/deformation  behav-­‐ iors   of   the   solid   electrolyte   interface   (SEI)   layers,   and   structural   and   morphological   changes.   A   comparison   study  on  two  different  sized  materials  (rod-­‐shape,  diame-­‐ ter  5  vs.  53  nm)  is  also  carried  out  and  preliminary  results   show  interesting  size-­‐dependent  properties,  that  the  5  nm   β-­‐FeOOH   has   more   serious   kinetic   restrictions,   and   thus   lower   capacities,   while   it   shows   better   cycling   stability.   Moreover,   while   the   origins   of   the   large   overpotential   of   electrode   materials   have   been   usually   considered   to   be   complicated   and   difficult   to   determine,60,   69   we   suggest,   based   on   our   study,   that   the   overpotential   should   be   mainly   due   to   the   features   of   the   surface   concentration   dependent   potential   and   the   slow   diffusion   of   Na+;   the   presence   of   the   SEI   layers   may   also   contribute   to   the   overpotential.     EXPERIMENTAL  SECTION   Synthesis.   A   simple   hydrolysis   method   was   used   to   con-­‐ trol  the  size  of  β-­‐FeOOH  nanorods.59,  70  For  the  synthesis   of   β-­‐FeOOH   nanorods   with   a   mean   diameter   of   53   nm   (see   Supporting   Information   Figure   S1   for   size   distribu-­‐ tion),   10   mL   of   0.5   M   aqueous   FeCl3   (Sigma-­‐Aldrich)   was  

Page 2 of 11

added   into   70   mL   of   H2O   in   a   100   mL   plastic   bottle   with   a   0.8  mm  needle  hole.  After  shaking  for  1–2  min,  the  bottle   was   kept   in   an   oven   at   100   °C   for   24   h.   The   precipitates   were   collected   after   cooling   down   and   washed   with   de-­‐ ionized  water  and  ethanol  several  times  and  finally  dried   at   80   °C   overnight.   For   the   synthesis   of   β-­‐FeOOH   nano-­‐ rods  with  a  mean  diameter  of  5  nm,  5  mmol  of  LiOH  (Alfa   Aesar)  was  added  in  the  beginning  while  all  other  condi-­‐ tions  were  kept  the  same.   Characterization.   XRD   analysis   was   carried   out   with   a   Shimadzu  thin-­‐film  diffractometer  using  Cu  Kα  radiation   for   both   the   prepared   materials   and   the   electrodes   after   discharge/charge.   Field   emission   scanning   electron   mi-­‐ croscopy  (FESEM)  images  were  collected  on  a  JEOL  JSM-­‐ 6340F  microscope.  High-­‐resolution  transmission  electron   microscopy  (HRTEM)  images  were  observed  using  a  JEOL   JEM   2010   microscope.   XANES   was   carried   out   at   Singa-­‐ pore   Synchrotron   Light   Source   (SSLS),   XAFCA   beamline.   For   the   XRD   and   XANES   analysis   on   the   material   after   discharge/charge,  all  the  electrodes  were  sealed  with  Kap-­‐ ton  tape  to  prevent  oxidation.   Electrochemical   Measurements.   The   active   material   β-­‐ FeOOH   nanorods   were   mixed   with   graphite   (Sigma-­‐ Aldrich),   conductive   carbon   (Super   P)   and   binder   (sodi-­‐ um  carboxymethyl  cellulose,  CMC,  Sigma-­‐Aldrich)  in  de-­‐ ionized   water   in   a   weight   ratio   of   50   :   30   :   5   :   15.   The   re-­‐ sultant  slurry  was  coated  onto  a  copper  current  collector   and  dried  at  80  °C  overnight  under  vacuum.  The  graphite   was  used  to  improve  stability  of  the  electrode.59  The  elec-­‐ trodes   were   assembled   with   Li   metal   as   counter   elec-­‐ trodes   in   a   2032   coin   cell   in   an   Ar-­‐filled   glove   box.   Glass   fiber   separators   (EL-­‐CELL,   ECC1-­‐01-­‐0012-­‐C)   were   used   to   keep  the  electrodes  apart.  The  electrolyte  was  1  M  NaClO4   (Sigma-­‐Aldrich)   in   propylene   carbonate   (Sigma-­‐Aldrich)   with   5   wt%   fluoroethylene   carbonate   (FEC,   Sigma-­‐ Aldrich).  Galvanostatic  discharge  and  charge  (0.005  –  2.8   V)   tests   were   performed   on   Battery   Testing   Equipment   (Neware  Electronic  Co.,  China)  at  different  current  densi-­‐ ties  under  ambient  temperature.  Cyclic  voltammetry  (CV)   and  electrochemical  impedance  spectroscopy  (EIS)  meas-­‐ urements   were   performed   on   a   Bio-­‐Logic   SP-­‐150   potenti-­‐ ostat.  The  frequency  range  and  AC  amplitude  used  for  the   EIS  measurements  were  1  M  Hz  –  0.01  Hz  and  10  mV,  re-­‐ spectively.     RESULTS  AND  DISCUSSION   The   β-­‐FeOOH   studied   here   was   synthesized   by   a   simple   hydrolysis   process.   They   own   rod-­‐like   shapes   with   con-­‐ trolled   sizes   (mean   diameters:   53   and   5   nm,   Figure   S1).   The  phase  structure  was  confirmed  by  XRD  analysis  (Fig-­‐ ure  S2).  The  specific  surface  areas  are  16  and  105  m2  g-­‐1  for   the  53  and  5  nm  materials,  respectively.70  We  first  investi-­‐ gate   the   sodium   storage   behaviors   of   the   material   based   on   β-­‐FeOOH   rods   with   mean   diameter   of   53   nm.   And   then  conduct  a  comparison  study  with  a  5  nm  β-­‐FeOOH.   Figure   1a   shows   the   first   five   discharge/charge   curves   of  the  53  nm  β-­‐FeOOH  at  a  current  density  of  80  mA  g-­‐1.   The  first  discharge  (sodium  insertion)  curve  has  one  short  

ACS Paragon Plus Environment

plateau  between  1.1  and  1.2  V  vs.  Na+/Na  at  the  beginning   of  the  sodium  insertion  process,  followed  by  a  falling  po-­‐ tential   till   the   end   of   discharge   at   a   potential   of   0.005   V   vs.  Na+/Na.  There  is  a  phase  transformation  process  as  the   structure   loses   long-­‐range   order   without   any   significant   XRD   peaks   after   insertion   of   sodium   (Figure   S3).   Upon   charge  (sodium  extraction),  the  voltage  profile  is  lack  of  a   well-­‐defined   plateau,   composed   of   only   sloping   regions,   indicating   the   poor   reversibility   of   the   reaction   occurred   at   the   plateau   during   the   initial   discharge.   The   phase   transformation  behavior  is  found  to  be  irreversible  as  the   material  remains  amorphous  after  charged  back  to  2.8  V.   The   discharge   and   charge   processes   deliver   a   capacity   of    

2.5 2.0 1.5 1.0 0.5 0.0 0

200 400 600 800 1000 Capacity (mAh g-1)

792   and   627   mAh   g-­‐1,   respectively,   corresponding   to   an   irreversible   loss   of   165   mAh   g-­‐1   and   a   CE   of   79%.   Such   a   high  reversible  capacity  is  close  to  the  theoretical  capacity   of   alloying   Sb   anode.   The   CE   is   at   least   10%   higher   than   those  reported  previously  for  iron-­‐based  materials.   62-­‐65,  67,   68   To   be   able   to   exclude   the   capacity   contribution   of   the   carbons   (Super   P   and   graphite)   which   were   used   to   im-­‐ prove  the  conductivity  and  stability  of  the  electrode,59  we   also   tested   their   sodium   storage   capacity,   and   found   a   very   low   reversible   capacity   of   ~20   mAh   g-­‐1   (Figure   S4),   which   is   negligible   compared   to   the   large   capacity   deliv-­‐ ered  from  β-­‐FeOOH.  

b

100 900

Discharge Charge

600

85

75 2

4 6 8 Cycle number

10

0.4 0.2

90

80

300 0

c

95

CE

CE (%)

1st 2nd 3rd 4th 5th

3.0

Capacity (mAh g-1)

a Potential (V, vs. Na+/Na)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Current (A g-1

Page 3 of 11

0.0 -0.2 -0.4 -0.6

sh Fre 1st 2nd 3rd 4th

0.0 0.5 1.0 1.5 2.0 2.5 3.0 + Potential (V vs. Na /Na)

Figure  1.  Electrochemical   properties  of   β-­‐FeOOH   (Dave  =  53  nm).  a)  Galvanostatic  discharge/charge  curves  at  a  current  density  of   80  mA  g-­‐1;  b)  first  10  cycle  capacity  and  CE;  c)  CV  curves  at  a  scanning  rate  of      0.2  mV  s-­‐1.  

In   the   second   cycle,   both   of   the   discharge   and   charge   profiles  are  quite  unlike  the  first.  There  is  also  an  absence   of  a  well-­‐defined  plateau  during  the  second  discharge  and   the  sloping  profile  features  a  significant  voltage  step  from   0.4  to  0.3  V.  During  charge,  the  profile  becomes  smoother   compared  with  the  first  cycle.  These  suggest  that  the  ma-­‐ terial   is   instable   after   one   cycle.   During   the   third   cycle,   the   discharge   profile   is   still   distinctively   different   from   the   second,   while   the   charge   profile   remains   nearly   the   same.  In  the  subsequent  cycles,  the  discharge  and  charge   profiles  of  each  cycle  are  almost  invariant,  and  only  shows   a  small  but  continuous  reduction  in  capacity.  Because  the   shape  of  the  curves  start  to  be  almost  the  same  from  the   second  charge  (i.e.  after  the  second  discharge),  the  mate-­‐ rial   become   relatively   stable   since.   The   discharge   and   charge  capacities  of  the  second  cycle  are  580  and  523  mAh   g-­‐1,  respectively.  The  capacity  decreases  gradually  after  the   second  cycle.  A  charge  capacity  of  351  mAh  g-­‐1  is  obtained   after   10   cycles   (Figure   1b).   The   CE   reaches   close   to   100%   after  a  few  cycles.  The  material  also  owns  a  good  rate  ca-­‐ pability,  at  a  current  of  500  mA  g-­‐1,  it  has  a  reversible  ca-­‐ pacity   of   509,   456   and   187   mAh   g-­‐1   for   the   1st,   2nd   and   100th  cycle,  respectively  (Figure  S5).   CV  was  also  employed  to  investigate  the  sodium  storage   behavior  of  β-­‐FeOOH.  Figure  1c  shows  CV  profiles  of  the   53  nm  β-­‐FeOOH.  Each  cycle  presents  several  peaks  during   both   reduction   and   oxidation,   characteristic   of   multiple-­‐ intermediate   processes   of   insertion   and   extraction.   Dur-­‐ ing   the   first   cycle,   the   material   show   reduction   peaks   at  

0.90,   0.59   and   near   0   V;   and   one   pronounced   oxidation   peak   at   1.39   V   associated   with   several   shoulder   peaks   on   both  sides.  During  the  second  cycle,  there  is  a  significant   reduction  in  intensity  of  the  main  peaks  because  the  reac-­‐ tions   are   partially   reversible.   The   reduction   peaks   are   al-­‐ most  at  the  same  positions  with  the  first  cycle,  while  the   main  oxidation  peaks  appear  at  different  positions  due  to   an   activation   process.   From   the   second   oxidation   cycle,   the   characteristics   of   the   CV   profile   remain   practically   identical,  with  only  small  reduction  in  peak  intensity  and   small  shifts  of  the  peak  positions,  indicating  the  material   becomes   more   stable,   consistent   with   the   discharge   and   charge   processes.   The   peak   at   0.90   V   can   be   assigned   to   the  reaction  at  the  short  plateau  during  the  first  discharge   corresponding   to   the   first   iron   reduction   step.   The   most   pronounced  reduction  peak  at  0.59  V  during  the  first  cy-­‐ cle   corresponding   to   the   second   iron   reduction   step.   It   reduces  fast  in  the  following  cycles,  and  nearly  vanishes  in   the  fourth  cycle.  The  reduction  current  near  0  V  is  from  a   combination   of   reduction   of   iron   and   formation   of   SEI   layers   as   confirmed   by   XANES   and   microscopy   studies   below.  The  oxidation  peaks  during  anodic  cycles  originate   from   deformation   of   the   SEI   layers   and   corresponding   multiple-­‐intermediate  sodium  extraction  reactions.   To   further   understand   the   redox   processes   associated   with   sodium   insertion   and   extraction,   we   carried   out   XANES   studies   on   the   electrodes   at   different   sodiation   (discharge/charge)   states   during   the   first   two   cycles.   Comparison  of  the  Fe  K-­‐edge  XANES  spectra  allows  us  to  

ACS Paragon Plus Environment

 

Chemistry of Materials acquire  valence  information  of  iron  in  the  electrode  as  the   overall  edge  position  shifts  to  lower  energies  upon  reduc-­‐ tion,   due   to   the   decrease   in   binding   energy   of   the   Fe   1s2   electrons.   The   spectra   are   shown   in   Figure   2a   –   f,   the   points   chosen   for   analysis   shown   in   Figure   2g,   and   the   corresponding   1st   derivative   plots   shown   in   Figure   S6.   During  the  first  discharge  (Figure  2a),  we  can  see  the  ab-­‐ sorption   edge   position   shifts   continuously   till   the   cut-­‐off   potential   of   0.005   V   vs.   Na+/Na.   This   means   the   reduction   of   iron   occurs   over   the   whole   potential   region.   It   is   ob-­‐ served   that   the   spectra   for   the   low   potential   electrodes   (0.3   and   0.005   V)   exhibit   a   significantly   higher   pre-­‐edge   jump   compared   with   those   at   high   potentials,   indicating   the  presence  of  metallic  iron  in  the  electrodes  at  low  po-­‐

e

Fe 1D, 0.005 V 1C, 0.55 V 1C, 2.0 V 1C, 2.8V β-FeOOH

7110 7125 7140 7155 Energy (eV)

f

1D, 0.005 V 2D, 0.005 V

7140

7200 7260 Energy (eV)

d

β-FeOOH 1C, 2.8 V 2D, 0.8 V 2D, 0.3 V 2D, 0.005 V Fe

7110 7125 7140 7155 Energy (eV)

g

1D

4.0

β-FeOOH 1D, 1.1 V 1C, 2.8 V 2C, 2.8 V

7110 7125 7140 7155 Energy (eV)

Normalized absorption (a.u.)

c Normalized absorption (a.u.)

7110 7125 7140 7155 Energy (eV)

tentials,  which  is  further  confirmed  on  the  extended  X-­‐ray   absorption   fine   structure   (EXAFS)   spectrum   with   the   presence   of   the   iron   peak   (Figure   S7).   However,   even   at   0.005   V,   the   absorption   edge   position   is   clearly   higher   than   that   of   the   iron   reference,   indicating   the   average   valence   of   iron   at   0.005   V   is   above   0.   These   results   demonstrate  the  formation  of  metallic  iron  at  low  poten-­‐ tials,  but  only  partial  of  the  iron  in  the  material  is  reduced   to  the  metallic  state,  which  is  consistent  with  that  the  1st   discharge  capacity  of  792  mAh  g-­‐1  is  lower  than  the  theo-­‐ retical  capacity  of  905  mAh  g-­‐1  from  Fe(III)  to  Fe(0).  This   partial   reduction   of   iron   is   different   from   its   Li   storage   behavior   where   the   iron   could   be   completely   reduced   to   its  lowest  valence.59  

Potential (V, vs. Na+/Na)

β-FeOOH 1D, 1.1 V 1D, 0.8 V 1D, 0.3 V 1D, 0.005V Fe

Normalized absorption (a.u.)

b

Normalized absorption (a.u.)

Normalized absorption (a.u.)

a

Normalized absorption (a.u.)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 11

Fe 2D, 0.005 V 2C, 0.55 V 2C, 2.0 V 2C, 2.8 V β-FeOOH

7110 7125 7140 7155 Energy (eV) 2D

1C

3.2

2.8

2.4

2.0

1.6

1.1

0.8

0.8 0.3

0.8

0.55 0.005

0.0 0

10

20 Time (h)

2C 2.8 2.0 0.55 0.3 0.005 30

 

Figure  2.   Fe   K-­‐edge   XANES   spectra   for   β-­‐FeOOH   (Dave   =   53   nm)   at   different   sodiation   states.   a)   Electrodes   after   the   first   discharge   to  1.1,  0.8,  0.3  and  0.005  V,  comparing  the  β-­‐FeOOH  and  metallic  Fe  as  references;  b)  electrodes  after  the  first  discharge  to  0.005  V   and  after  the  first  charge  to  0.55,  2.0  and  2.8  V,  and  references;  c)  electrodes  after  the  first  charge  to  2.8  V  and  after  the  second   discharge  to  0.8,  0.3,  and  0.005  V,  and  references;   d)  electrodes  after  the  second  discharge  to  0.005  V  and  after  the  second  charge   to    0.55,  2.0  and  2.8  V,  and  references;  e)  electrodes  after  the  first  and  second  charge  to  0.005  V;  f)  electrodes  after  the  first  dis-­‐ charge  to  1.1  and  after  the  first  and  second  charge  to  2.8  V,  and  β-­‐FeOOH  reference;  g)  points  chosen  for  XANES  analysis.  1D,  1C,   2D,  and  2C  in  the  legends  stand  for  first  discharge,  first  charge,  second  discharge,  and  second  charge,  respectively.  

Upon   subsequent   charge   (Figure   2b),   oxidation   of   iron   starts   above   0.55   V   as   the   plots   almost   overlaps   at   0.005   and   0.55   V.   At   the   final   charge   potential   of   2.8   V,   iron   is   still   not   fully   oxidized   to   Fe(III)   as   the   absorption   edge   position  is  located  slightly  lower  than  the  one  for  the  pris-­‐ tine   β-­‐FeOOH.   This   results   in   an   irreversible   loss   of   the   capacity   consistent   with   the   galvanostatic   cycling   meas-­‐ urement.   To   determine   the   oxidation   state   at   2.8   V   after   one  cycle,  we  compare  the  spectrum  at  this  potential  with   the   one   at   1.1   V   during   the   first   charge,   and   find   the   ab-­‐ sorption   edge   position   at   2.8   V   is   higher   than   the   one   at   1.1   V.   The   valence   of   iron   at   1.1   V   is   +2.6   calculated   from   the   capacity   of   128   mAh   g-­‐1   at   this   potential,   assuming  

there   is   no   formation   of   SEI   layers   at   this   relatively   high   potential  and  all  the  capacity  is  delivered  from  the  reduc-­‐ tion  of  iron.  Thus,  the  valence  of  iron  at  2.8  V  is  between   +2.6   and   +3.   The   irreversible   loss   of   the   capacity   caused   by   the   incomplete   oxidation   of   iron   should   be   less   than   128   mAh   g-­‐1,   lower   than   the   total   loss   of   165   mAh   g-­‐1   in   the   first   cycle,   indicating   there   are   other   irreversible   (or   par-­‐ tially   irreversible)   reactions,   which   can   be   attributed   to   the   formation/deformation   of   SEI   layers   as   confirmed   below   by   field   emission   scanning   electron   microscopy   (FESEM)   and   high-­‐resolution   transmission   electron   mi-­‐ croscopy  (HRTEM).  

ACS Paragon Plus Environment

Page 5 of 11

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

The   spectra   for   the   second   cycle   show   the   same   trend   (Figure   2c   and   d).   Upon   discharge,   iron   is   gradually   re-­‐ duced  until  0.005  V;  and  its  oxidation  starts  above  0.55  V   during   charge.   The   oxidation   state   at   0.005   V   is   very   simi-­‐ lar  to  the  one  after  the  first  discharge  with  their  edge  po-­‐ sition   located   at   quite   the   same   position   (Figure   2e   and   Figure   S6e).   However,   we   can   find   the   two   spectra   have   different  shapes,  indicating  different  local  chemical  envi-­‐ ronments  of  iron.  The  oxidation  state  of  iron  at  2.8  V  after   two  cycles  is  close  to  the  one  at  1.1  V  during  the  first  dis-­‐ charge  (Figure  2f),  lower  than  that  after  the  first  charge  at   2.8  V,  in  agreement  with  the  capacity  reduction  observed.   The  morphological  changes  upon  sodium  insertion  and   extraction  as  well  as  the  formation/deformation  behaviors   of   SEI   layers   were   studied   by   both   FESEM   and   HRTEM.   Figure   3a   and   Figure   S8a   show   the   FESEM   images   of   fresh  electrode.  The  original  rod-­‐like  shape  of  the  materi-­‐ al  remains  in  the  electrode.  We  can  also  find  small  Super   P   carbon   and   micro-­‐sized   graphite   used   in   the   electrode   in  the  FESEM  images.  After  insertion  of  a  small  amount  of   sodium  at  1.1  V,  the  material  starts  to  crack  into  small  na-­‐ noparticles   (Figure   3b),   indicating   large   volume   strains   upon  sodium  insertion.  After  discharged  to  0.8  V,  there  is   further   cracking   of   the   particles   which   become   signifi-­‐ cantly  smaller  than  at  1.1  V  (Figure  3c).  Discharge  to  0.3  V   doesn’t  lead  to  an  obvious  size  change  of  small  nanoparti-­‐ cles  (Figure  3d).  However,  at  this  potential,  we  can  easily   find   large   microsized   particles   (indicated   by   the   ellipse)    

with   morphology   similar   to   the   final   discharge   state   at   0.005  V,  which  seem  to  be  agglomerates  from  small  parti-­‐ cles.   At   0.005   V,   microsized   particles   are   the   domain   characteristic   of   the   electrode.   There   are   also   a   lot   of   small   nanoparticles   on   the   surface   (see   also   Figure   S8b).   We   attribute   the   appearance   of   the   microsized   particles   observed  at  low  potentials  to  the  formation  of  SEI  layers.   The   SEI   layers   form   on   the   surface   of   the   small   particles   and  make  them  interconnect  with  each  other.  If  the  layers   are   thick   enough,   the   boundaries   between   nanoparticles   disappear  and  they  grow  into  large  particles.  The  SEI  lay-­‐ ers  are  formed  inhomogeneously,  evidenced  by  the  coex-­‐ istence   of   the   nano-­‐   and   micro-­‐particles   at   the   low   poten-­‐ tials.  After  charged  back  to  2.8  V,  the  majority  of  the  SEI   layers   deforms,   and   the   material   remains   to   be   small   nano-­‐spheres  instead  of  the  original  rod  shape  with  larger   sizes.  The  cracking  of  the  original  particle  and  formation   of   the   SEI   layers   are   confirmed   by   HRTEM   analysis.   At   0.005   V   (Figure   3g),   particles   with   size   of   ~100   nm   are   observed   in   SEI   matrix.   Interestingly,   these   particles   are   composed   of   very   small   nanoparticles.   This   is   similar   to   conversion  anodes  after  lithiation.60  At  2.8  V  (Figure  3h),   we  can  still  find  a  thin  SEI  layer  on  the  particles,  indicat-­‐ ing  the  deformation  of  SEI  layer  is  partially  reversible.  The   extreme   volume   strains   observed   here   has   the   serious   drawback   that   causes   instability   of   the   electrode   leading   to  poor  cyclability  (Figure  1b  and  Figure  S5).  

  Figure   3.  SEM  and  TEM  images  of  β-­‐FeOOH  (Dave  =  53  nm)  at  different  sodiation  states.  SEM  images  of  a)  fresh  electrode;  b,  c,  d   and  e)  electrodes  after  the  first  discharge  to  1.1,  0.8,  0.3  and  0.005,  respectively;  f)  electrode  after  the  first  charge  to  2.8  V.  TEM   images  of  the  electrodes  after  the  first  discharge  to  0.005  V  (g);  and  after  the  first  charge  to  2.8  V  (h).  The  ellipse  in  (d)  indicates   similar  morphology  observed  in  (e).  The  arrows  in  (g)  and  (h)  indicate  the  SEI  layers.  

Finally,   we   compare   the   sodium   storage   performance   of   the  53  nm  material  with  the  5  nm  one.  Figure  4  shows  the   electrochemical   properties   of   the   5   nm   β-­‐FeOOH.   The   materials  has  a  relatively  low  capacity  and  rate  capability  

(495  and  266  mAh  g-­‐1  at  80  and  500  mA  g-­‐1),  while  exhibits   rather   stable   cycling   behavior.   After   two   cycles,   the   dis-­‐ charge/charge  curves  overlap  at  a  current  of  80  mA  g-­‐1  and   the   reversible   capacity   remains   almost   the   same   at   each  

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

termine   the   origins   of   overpotential,60,   69   it   seems   clear   here   that   the   overpotential   of   the   β-­‐FeOOH   electrode   is   at   least   mainly   derived   from   the   features   of   the   surface   concentration  dependent  potential  and  the  slow  diffusion   of  Na+.    

rate.   The   improved   cycling   performance   is   probably   be-­‐ cause   that   smaller   particles   provide   possibility   of   better   contacts   between   electrode   material,   conductive   carbon   and  binder  which  is  beneficial  for  the  maintenance  of  the   stability  of  the  electrode,  provided  that  there  is  sufficient   conductive  carbon  and  binder  used.  

Comparing   the   two   materials   in   the   relaxation   meas-­‐ urements,   we   can   find   that   Na+   diffuses   slower   in   the   5   nm   material   as   it   takes   longer   to   achieve   the   quasi-­‐ equilibrium   state   and   the   5   nm   material   also   has   a   large   overpotential,   for   both   the   fresh   electrodes   and   the   ones   after   three   cycles.   It’s   interesting   that   the   smaller-­‐sized   material   needs   longer   time   to   achieve   the   quasi-­‐ equilibrium   state   because   it   has   a   shorter   distance   for   Na+   to   diffuse   to   the   very   inner   part.   The   slower   diffusion   of   the  5  nm  material  should  lead  to  the  larger  overpotential   observed.  Understanding  this  difference  between  the  two   materials   should   be   helpful   in   designing   materials   that   can   avoid   the   large   overpotential.   This   would   be   worth   further   research   efforts.   When   comparing   the   measure-­‐ ments  on  the  fresh  electrodes  and  on  the  ones  after  three   cycles,   we   can   find   the   electrodes   after   three   cycles   need   longer   time   to   achieve   quasi-­‐equilibrium   states   for   both   two   materials,   indicating   slower   diffusion   kinetics   after   cycling.  This  might  be  attributed  to  the  slow  diffusion  of   Na+   in   the   SEI   matrix   that   forms   at   low   potentials   and   remains   still   partially   at   high   potentials   as   discussed   above.  In  this  regard,  building  advanced  SEI  layers  where   Na+   can   transfer   fast   will   be   favorable   for   reducing   the   overpotential.   This   might   be   achieved   by   changing   elec-­‐ trolyte   and/or   adding   additives.   It   is   also   found   that   the   overpotentials   obtained   using   the   two   electrodes   after   three  cycles  are  larger  than  the  ones  before  cycling,  indi-­‐ cating  that  the  presence  of  SEI  contributes  to  the  overpo-­‐ tential.  

Potential   relaxation   measurements   (Figure   5)   per-­‐ formed  on  the  two  β-­‐FeOOH  provide  further  insights  into   their  sodium  storage  behaviors.    Two  electrodes  at  differ-­‐ ent   sodiation   states   were   used   for   the   measurements   for   both  materials,  fresh  electrodes  (Figure  5a)  and  electrodes   after  three  cycles  (Figure  5b).  Sodium  was  inserted  step  by   step  using  a  very  low  current  of  20  mA  g-­‐1  for  1  h  and  wait-­‐ ing  after  each  step  for  10  h  until  a  capacity  of  100  mAh  g-­‐1   was   achieved.   Afterwards,   the   batteries   were   left   until   stable,   i.e.,   achieving   their   quasi-­‐equilibrium   state.   The   quasi-­‐equilibrium   potentials   after   insertion   of   a   capacity   of  100  mAh  g-­‐1  are  2.64  and  2.61  V  for  the  fresh  electrodes   with   the   53   and   5   nm   materials,   respectively.   Materials   with   different   sizes   have   different   specific   surface   areas   and  they  are  in  some  cases  considered  to  affect  the  equi-­‐ librium  potential  of  electrode  materials  due  to  the  surface   energy   contribution.60,   71   Here,   the   slight   difference,   0.03   V,  of  the  potentials  indicates  that  the  surface  energy  con-­‐ tribution  has  limited  effects  on  the  equilibrium  potential   for   β-­‐FeOOH   for   sodium   storage.   When   using   the   elec-­‐ trodes  after  three  cycles,  the  quasi-­‐equilibrium  potentials   are  2.68  and  2.60  V  for  the  53  and  5  nm  materials,  respec-­‐ tively.  Apart  from  the  possible  size  difference  of  these  two   materials  after  cracking,  the  enlarged  potential  difference,   0.08  V,  may  be  also  due  to  the  different  sodiation  states  of   the   two   materials,   because   they   may   contain   different   amount  of  Na+  after  three  cycles.     In   all   the   cases,   it   takes   very   long   for   the   material   to   achieve   the   quasi-­‐equilibrium   state,   more   than   300   h,   being   the   fastest,   for   the   53   nm   material   initially   in   the   fresh  electrode,  and  more  than  1000  h  or  even  more  than   3000  h  for  the  others.  This  suggests  a  very  slow  Na+  diffu-­‐ sion  rate  in  the  material.  The  overpotentials,  as  indicated   in  Figure  5,  are  also  found  to  be  very  large,  >1  V,  for  all  the   cases.   The   slow   diffusion   and   the   large   overpotentials   suggest   that   the   potential   of   the   material   highly   depend   on   the   concentration   of   Na+   in   the   surface   part   of   the   ma-­‐ terial  (similar  to  a  pseudo-­‐capacitive  behavior72).  With  the   decrease  of  the  surface  concentration  via  diffusion  of  Na+   toward   inner   where   its   concentration   is   lower,   the   open   circuit   potential   increases   until   the   equilibrium   state   is   reached.  While  it  is  usually  thought  to  be  difficult  to  de-­‐

Page 6 of 11

As  mentioned,  the  5  nm  material  has  slower  diffusion  of   Na+.  It  suffers  from  more  serious  kinetic  restrictions  dur-­‐ ing   sodium   insertion   and   extraction,   and   thus   leads   to   a   lower  capacity  and  rate  capability.  One  may  also  consider   other   kinetic   factors   for   the   relatively   low   capacity   and   rate  capability  of  the  5  nm  material,  such  as  the  conduc-­‐ tivity   and   the   interfacial   resistance.   However,   electro-­‐ chemical   impedance   spectroscopy   analysis   (Figure   S9)   has   shown   the   two   batteries   have   the   same   ohmic   re-­‐ sistance  and  the  5  nm  material  has  smaller  interfacial  re-­‐ sistances  than  the  53  nm  one.    

 

ACS Paragon Plus Environment

0.5

0m A g -1 20

00

300

mA g -1

0m A g -1

1.0

10

1.5

600

50

2.0

Discharge Charge

mA g -1 0m A g -1

2.5

b

80

1st 2nd 3rd 4th 5th

3.0

20

Potential (V, vs. Na+/Na)

a

0.0 0

200 400 600 Capacity (mAh g-1)

800

0

10

20 30 40 Cycle number

50

 

Figure   4.  Electrochemical  properties  of  β-­‐FeOOH  (Dave  =  5  nm).  a)  Galvanostatic  discharge/charge  curves  at  a  current  density  of  80   mA  g-­‐1;  b)  cycling  performance  and  rate  capability.  

Potential (V, vs. Na+/Na)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Capacity (mAh g-1)

Page 7 of 11

3.0 a) 2.64 V 2.61 V 2.5 2.0 1.32 V 1.44 V 1.5 1.0 0 500 1000 1500 Time (h) 3.0 b) 2.5 2.0 1.5 1.0 0

-1

100 mAh g 53 nm 5 nm 2000

0

2.68 V 2.60 V 1.66 V 1500

1.67 V

3000 Time (h)

Rest

20 40 Time (h)

60

Rest 100 mAh g-1 Rest 53 nm 5 nm

4500

6000 0

40 80 Time (h)

Figure  5.  Time  for  achieving  quasi-­‐equilibrium  states  after  insertion  of  sodium  capacity  of  100  mAh  g-­‐1.  Measurements  performed   on  a)  fresh  electrode  and  b)  electrodes  after  three  cycles  and  when  the  OCVs  became  relatively  stable.  Sodium  was  inserted  step  by   step  using  a  very  low  current  of  20  mA  g-­‐1  for  1  h  and  waiting  after  each  step  for  10  h  until  a  capacity  of  100  mAh  g-­‐1  was  achieved.   Afterwards,   the   batteries   were   left   until   quasi-­‐equilibrium.   The   quasi-­‐equilibrium   potentials   and   the   overpotentials   after   insertion   of  a  capacity  of  100  mAh  g-­‐1  of  Na+  have  been  indicated.    

CONCLUSIONS   Earth-­‐abundant   β-­‐FeOOH   has   been   investigated   as   a   novel   anode   for   NIBs.   The   53   nm   sample   exhibits   a   high   reversible   capacity   (627   and   523   mAh   g-­‐1   for   the   first   and   second  cycle,  respectively,  at  a  current  of  80  mAh  g-­‐1.  Alt-­‐ hough  it  also  shares  the  same  drawback,  fast  capacity  fad-­‐ ing,   of   other   high-­‐capacity   anodes,   the   material   holds   great  promise  for  sustainable  NIB  applications.     A   range   of   techniques   including   CV,   XANES,   XRD   and   microscopy   were   used   to   understand   the   dis-­‐ charge/charge  processes.  The  material  has  a  phase  trans-­‐ formation  from  crystalline  to  amorphous  after  insertion  of   sodium.   The   phase   transformation   behavior   is   irreversi-­‐ ble.  The  sodium  insertion  process  is  accompanied  by  ob-­‐ vious  particle  cracking.  The  conversion  of  the  material  to   metallic  iron  at  low  potentials  is  demonstrated.  However,   the   conversion   process   is   incomplete   even   at   a   cut-­‐off   potential  of  0.005  V  vs.  Na+/Na,  only  partial  of  the  materi-­‐

al  is  converted  to  metallic  iron.  At  a  charge  cut-­‐off  poten-­‐ tial   of   2.8   V   vs.   Na+/Na,   iron   can   be   oxidized   up   to   a   va-­‐ lence   between   +2.6   and   +3,   instead   of   back   to   +3.   This   contributes   to   the   irreversible   loss   of   the   capacity.   The   formation   of   the   SEI   layers   at   low   potentials   was   also   demonstrated.  Its  deformation  occurs  during  charge  pro-­‐ cess,  but  still  incomplete  at  2.8  V,  which  also  contributes   to  the  irreversible  capacity.     A  comparison  of  sodium  storage  properties  on  two  dif-­‐ ferent  sized  materials  (5  vs.  53  nm)  shows  interesting  size-­‐ dependent  properties,  that  the  5  nm  β-­‐FeOOH  has  more   serious   kinetic   restrictions,   and   thus   lower   capacities,   while   it   shows   better   cycling   stability.   Origins   of   the   large   overpetential  have  been  suggested  to  be  mainly  due  to  the   features  of  the  surface  concentration  dependent  potential   and  the  slow  diffusion  of  Na+;  and  the  presence  of  the  SEI   layers  may  also  contribute  to  the  overpotential.    

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ASSOCIATED  CONTENT     Supporting  Information.     HRTEM   images   of   the   prepared   materials,   XRD   patterns   of   the   prepared   β-­‐FeOOH   and   the   materials   after   dis-­‐ charge/charge,   cycling   performance   of   the   53   material   at   a   -­‐1 current   of   500   mAh   g ,   1st   derivative   plots   of   the   Fe   K-­‐edge   XANES   spectra   in   Figure   2A,   EXAFS   spectra   of   pristine   β-­‐ FeOOH,   Fe   reference   and   the   electrode   after   the   first   dis-­‐ charge,    FESEM  images  of  fresh  electrode  and  electrode  after   discharge/charge,   capacity   data   for   graphite   and   Super   P   carbon   electrode,   and   EIS   spectra   of   electrodes   at   different   sodiation   states.   This   material   is   available   free   of   charge   via   the  Internet  at  http://pubs.acs.org.  

AUTHOR  INFORMATION   Corresponding  Author   [email protected];  [email protected]­‐star.edu.sg    

Funding  Sources   This  work  was  supported  by  MOE  Tier  1  Grants  (RGT8/13,   RG13/13,  and  RG131/14)  of  Singapore,  and  the  Singapore  Na-­‐ tional  Research  Foundation  under  its  Campus  for  Research   Excellence  And  Technological  Enterprise  (CREATE)  pro-­‐ gramme.  P.Y.  is  supported  by  SSLS  via  NUS  Core  Support  C-­‐ 380-­‐003-­‐003-­‐001.  

ACKNOWLEDGMENT     Authors   thank   the   Facility   for   Analysis,   Characterisation,   Testing   and   Simulation   (FACTS)   in   Nanyang   Technological   University   for   materials   characterizations.   Authors   also   thank   the   facility   support   provided   by   the   Singapore   Syn-­‐ chrotron  Light  Source.  

REFERENCES   (1)   Slater,  M.  D.;  Kim,  D.;  Lee,  E.;  Johnson,  C.  S.,  Sodium-­‐ Ion  Batteries.  Adv.  Funct.  Mater.  2013,  23,  947-­‐958.   (2)   Pan,   H.   L.;   Hu,   Y.   S.;   Chen,   L.   Q.,   Room-­‐Temperature   Stationary   Sodium-­‐Ion   Batteries   for   Large-­‐Scale   Electric   Energy   Storage.  Energy  Environ.  Sci.  2013,  6,  2338-­‐2360.   (3)   Yu,  H.;  Ren,  Y.;  Xiao,  D.;  Guo,  S.;  Zhu,  Y.;  Qian,  Y.;  Gu,   L.;   Zhou,   H.,   An   Ultrastable   Anode   for   Long-­‐Life   Room-­‐ Temperature  Sodium-­‐Ion  Batteries.  Angew.   Chem.   Int.   Ed.   2014,   53,  8963-­‐8969.   (4)   Wang,  S.;  Wang,  L.;  Zhu,  Z.;  Hu,  Z.;  Zhao,  Q.;  Chen,  J.,   All   Organic   Sodium-­‐Ion   Batteries   with   Na4C8H2O6.   Angew.   Chem.  Int.  Ed.  2014,  53,  5892-­‐5896.   (5)   Billaud,   J.;   Clément,   R.   J.;   Armstrong,   A.   R.;   Canales-­‐ Vázquez,   J.;   Rozier,   P.;   Grey,   C.   P.;   Bruce,   P.   G.,   β-­‐NaMnO2:   A   High-­‐Performance   Cathode   for   Sodium-­‐Ion   Batteries.   J.   Am.   Chem.  Soc.  2014,  136,  17243-­‐17248.   (6)   Hong,  S.  Y.;  Kim,  Y.;  Park,  Y.;  Choi,  A.;  Choi,  N.  S.;  Lee,   K.   T.,   Charge   Carriers   in   Rechargeable   Batteries:   Na   Ions   vs.   Li   Ions.  Energy  Environ.  Sci.  2013,  6,  2067-­‐2081.   (7)   Zhou,  T.;  Pang,  W.  K.;  Zhang,  C.;  Yang,  J.;  Chen,  Z.;  Liu,   H.   K.;   Guo,   Z.,   Enhanced   Sodium-­‐Ion   Battery   Performance   by   Structural   Phase   Transition   from   Two-­‐Dimensional   Hexagonal-­‐ SnS2  to  Orthorhombic-­‐SnS.  ACS  Nano  2014.   (8)   Kim,   H.;   Hong,   J.;   Park,   Y.-­‐U.;   Kim,   J.;   Hwang,   I.;   Kang,   K.,   Sodium   Storage   Behavior   in   Natural   Graphite   using   Ether-­‐ based  Electrolyte  Systems.  Adv.  Funct.  Mater.  2015,  25,  534-­‐541.   (9)   David,   L.;   Singh,   G.,   Reduced   Graphene   Oxide   Paper   Electrode:   Opposing   Effect   of   Thermal   Annealing   on   Li   and   Na   Cyclability.  J.  Phys.  Chem.  C  2014,  118,  28401-­‐28408.  

Page 8 of 11

(10)   Li,  Y.;  Xu,  S.;  Wu,  X.;  Yu,  J.;  Wang,  Y.;  Hu,  Y.-­‐S.;  Li,  H.;   Chen,   L.;   Huang,   X.,   Amorphous   Monodispersed   Hard   Carbon   Micro-­‐Spherules   Derived   from   Biomass   as   a   High   Performance   Negative   Electrode   Material   for   Sodium-­‐Ion   Batteries.   J.   Mater.   Chem.  A  2015,  3,  71-­‐77.   (11)   Xia,  X.;  Dahn,  J.  R.,  Study  of  the  Reactivity  of  Na/Hard   Carbon  with  Different  Solvents  and  Electrolytes.  J.   Electrochem.   Soc.  2012,  159,  A515-­‐A519.   (12)   Tang,   K.;   Fu,   L.   J.;   White,   R.   J.;   Yu,   L.   H.;   Titirici,   M.-­‐ M.;   Antonietti,   M.;   Maier,   J.,   Hollow   Carbon   Nanospheres   with   Superior  Rate  Capability  for  Sodium-­‐Based  Batteries.  Adv.   Energy   Mater.  2012,  2,  873-­‐877.   (13)   Komaba,   S.;   Ishikawa,   T.;   Yabuuchi,   N.;   Murata,   W.;   Ito,   A.;   Ohsawa,   Y.,   Fluorinated   Ethylene   Carbonate   as   Electrolyte   Additive   for   Rechargeable   Na   Batteries.   ACS   Appl.   Mater.  Interfaces  2011,  3,  4165-­‐4168.   (14)   Stevens,  D.  A.;  Dahn,  J.  R.,  The  Mechanisms  of  Lithium   and   Sodium   Insertion   in   Carbon   Materials.   J.   Electrochem.   Soc.   2001,  148,  A803-­‐A811.   (15)   Senguttuvan,   P.;   Rousse,   G.;   Seznec,   V.;   Tarascon,   J.-­‐ M.;   Palacin,   M.   R.,   Na2Ti3O7:   Lowest   Voltage   Ever   Reported   Oxide   Insertion   Electrode   for   Sodium   Ion   Batteries.   Chem.   Mater.  2011,  23,  4109-­‐4111.   (16)   Zou,  W.;  Li,  J.;  Deng,  Q.;  Xue,  J.;  Dai,  X.;  Zhou,  A.;  Li,  J.,   Microspherical   Na2Ti3O7   Prepared   by   Spray-­‐Drying   Method   as   Anode   Material   for   Sodium-­‐Ion   Battery.   Solid   State   Ionics   2014,   262,  192-­‐196.   (17)   Rousse,   G.;   Arroyo-­‐de   Dompablo,   M.   E.;   Senguttuvan,   P.;  Ponrouch,  A.;  Tarascon,  J.-­‐M.;  Palacín,  M.  R.,  Rationalization   of  Intercalation  Potential  and  Redox  Mechanism  for  A2Ti3O7  (A  =   Li,  Na).  Chem.  Mater.  2013,  25,  4946-­‐4956.   (18)   Pan,   H.;   Lu,   X.;   Yu,   X.;   Hu,   Y.-­‐S.;   Li,   H.;   Yang,   X.-­‐Q.;   Chen,   L.,   Sodium   Storage   and   Transport   Properties   in   Layered   Na2Ti3O7   for   Room-­‐Temperature   Sodium-­‐Ion   Batteries.   Adv.   Energy  Mater.  2013,  3,  1186-­‐1194.   (19)   Yu,   X.;   Pan,   H.;   Wan,   W.;   Ma,   C.;   Bai,   J.;   Meng,   Q.;   Ehrlich,  S.  N.;  Hu,  Y.-­‐S.;  Yang,  X.-­‐Q.,  A  Size-­‐Dependent  Sodium   Storage   Mechanism   in   Li4Ti5O12   Investigated   by   a   Novel   Characterization  Technique  Combining  in  Situ  X-­‐ray  Diffraction   and  Chemical  Sodiation.  Nano  Lett.  2013,  13,  4721-­‐4727.   (20)   Sun,  Y.;  Zhao,  L.;  Pan,  H.;  Lu,  X.;  Gu,  L.;  Hu,  Y.-­‐S.;  Li,   H.;  Armand,  M.;  Ikuhara,  Y.;  Chen,  L.;  Huang,  X.,  Direct  Atomic-­‐ Scale   Confirmation   of   Three-­‐Phase   Storage   Mechanism   in   Li4Ti5O12   Anodes   for   Room-­‐Temperature   Sodium-­‐Ion   Batteries.   Nat.  Commun.  2013,  4,  1870.   (21)   Liu,   J.;   Tang,   K.;   Song,   K.;   van   Aken,   P.   A.;   Yu,   Y.;   Maier,   J.,   Tiny   Li4Ti5O12   Nanoparticles   Embedded   in   Carbon   Nanofibers  as  High-­‐Capacity  and  Long-­‐Life  Anode  Materials  for   Both  Li-­‐Ion  and  Na-­‐Ion  Batteries.  Phys.  Chem.  Chem.  Phys.  2013,   15,  20813-­‐20818.   (22)   Zhao,   L.;   Pan,   H.-­‐L.;   Hu,   Y.-­‐S.;   Li,   H.;   Chen,   L.-­‐Q.,   Spinel   Lithium   Titanate   (Li4Ti5O12)   as   Novel   Anode   Material   for   Room-­‐Temperature   Sodium-­‐Ion   Battery.   Chinese   Phys.   B   2012,   21,  028201.   (23)   Yu,   P.;   Li,   C.;   Guo,   X.,   Sodium   Storage   and   Pseudocapacitive   Charge   in   Textured   Li4Ti5O12   Thin   Films.   J.   Phys.  Chem.  C  2014,  118,  10616-­‐10624.   (24)   Wang,   Y.;   Yu,   X.;   Xu,   S.;   Bai,   J.;   Xiao,   R.;   Hu,   Y.-­‐S.;   Li,   H.;   Yang,   X.-­‐Q.;   Chen,   L.;   Huang,   X.,   A   Zero-­‐Strain   Layered   Metal  Oxide  as  the  Negative  Electrode  for  Long-­‐Life  Sodium-­‐Ion   Batteries.  Nat.  Commun.  2013,  4,  2365.   (25)   Wu,   L.;   Buchholz,   D.;   Bresser,   D.;   Gomes   Chagas,   L.;   Passerini,   S.,   Anatase   TiO2   Nanoparticles   for   High   Power   Sodium-­‐Ion  Anodes.  J.  Power  Sources  2014,  251,  379-­‐385.   (26)   Wu,  L.;  Bresser,  D.;  Buchholz,  D.;  Giffin,  G.;  Castro,  C.   R.;  Ochel,  A.;  Passerini,  S.,  Unfolding  the  Mechanism  of  Sodium  

ACS Paragon Plus Environment

Page 9 of 11

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Insertion   in   Anatase   TiO2   Nanoparticles.   Adv.   Energy   Mater.   2014,  5,  1401142.   (27)   Xu,  Y.;  Lotfabad,  E.  M.;  Wang,  H.  L.;  Farbod,  B.;  Xu,  Z.   W.;  Kohandehghan,  A.;  Mitlin,  D.,  Nanocrystalline  Anatase  TiO2:   A   New   Anode   Material   for   Rechargeable   Sodium   Ion   Batteries.   Chem.  Commun.  2013,  49,  8973-­‐8975.   (28)   Xiong,   H.;   Slater,   M.   D.;   Balasubramanian,   M.;   Johnson,   C.   S.;   Rajh,   T.,   Amorphous   TiO2   Nanotube   Anode   for   Rechargeable  Sodium  Ion  Batteries.  J.   Phys.   Chem.   Lett.   2011,  2,   2560-­‐2565.   (29)   Li,   Z.;   Young,   D.;   Xiang,   K.;   Carter,   W.   C.;   Chiang,   Y.   M.,   Towards   High   Power   High   Energy   Aqueous   Sodium-­‐Ion   Batteries:   The   NaTi2(PO4)3/Na0.44MnO2   System.   Adv.   Energy   Mater.  2013,  3,  290-­‐294.   (30)   Sun   Il,   P.;   Gocheva,   I.;   Okada,   S.;   Yamaki,   J.   i.,   Electrochemical   Properties   of   NaTi2(PO4)3   Anode   for   Rechargeable  Aqueous  Sodium-­‐Ion  Batteries.  J.   Electrochem.   Soc.   2011,  158,  A1067-­‐A1070.   (31)   Wu,  X.  Y.;  Cao,  Y.  L.;  Ai,  X.  P.;  Qian,  J.  F.;  Yang,  H.  X.,   A  Low-­‐Cost  and  Environmentally  Benign  Aqueous  Rechargeable   Sodium-­‐Ion   Battery   Based   on   NaTi2(PO4)3-­‐Na2NiFe(CN)6   Intercalation  Chemistry.  Electrochem.  Commun.  2013,  31,  145-­‐148.   (32)   Komaba,   S.;   Matsuura,   Y.;   Ishikawa,   T.;   Yabuuchi,   N.;   Murata,   W.;   Kuze,   S.,   Redox   Reaction   of   Sn-­‐Polyacrylate   Electrodes   in   Aprotic   Na   Cell.   Electrochem.   Commun.   2012,   21,   65-­‐68.   (33)   Lin,   Y.-­‐M.;   Abel,   P.   R.;   Gupta,   A.;   Goodenough,   J.   B.;   Heller,   A.;   Mullins,   C.   B.,   Sn–Cu   Nanocomposite   Anodes   for   Rechargeable  Sodium-­‐Ion  Batteries.  ACS   Appl.   Mater.   Interfaces   2013,  5,  8273-­‐8277.   (34)   Liu,   Y.;   Xu,   Y.;   Zhu,   Y.;   Culver,   J.   N.;   Lundgren,   C.   A.;   Xu,   K.;   Wang,   C.,   Tin-­‐Coated   Viral   Nanoforests   as   Sodium-­‐Ion   Battery  Anodes.  ACS  Nano  2013,  7,  3627-­‐34.   (35)   Oh,   S.-­‐M.;   Myung,   S.-­‐T.;   Jang,   M.-­‐W.;   Scrosati,   B.;   Hassoun,   J.;   Sun,   Y.-­‐K.,   An   Advanced   Sodium-­‐Ion   Rechargeable   Battery   Based   on   a   Tin-­‐Carbon   Anode   and   a   Layered   Oxide   Framework   Cathode.   Phys.   Chem.   Chem.   Phys.   2013,   15,   3827-­‐ 3833.   (36)   Darwiche,   A.;   Marino,   C.;   Sougrati,   M.   T.;   Fraisse,   B.;   Stievano,   L.;   Monconduit,   L.,   Better   Cycling   Performances   of   Bulk   Sb   in   Na-­‐Ion   Batteries   Compared   to   Li-­‐Ion   Systems:   An   Unexpected  Electrochemical  Mechanism.  J.  Am.  Chem.  Soc.  2012,   134,  20805-­‐20811.   (37)   Qian,   J.;   Chen,   Y.;   Wu,   L.;   Cao,   Y.;   Ai,   X.;   Yang,   H.,   High   Capacity   Na-­‐Storage   and   Superior   Cyclability   of   Nanocomposite   Sb/C   Anode   for   Na-­‐Ion   Batteries.   Chem.   Commun.  2012,  48,  7070-­‐7072.   (38)   Baggetto,   L.;   Ganesh,   P.;   Sun,   C.-­‐N.;   Meisner,   R.   A.;   Zawodzinski,   T.   A.;   Veith,   G.   M.,   Intrinsic   Thermodynamic   and   Kinetic   Properties   of   Sb   Electrodes   for   Li-­‐Ion   and   Na-­‐Ion   Batteries:   Experiment   and   Theory.   J.   Mater.   Chem.   A   2013,   1,   7985-­‐7994.   (39)   Zhu,  Y.;  Han,  X.;  Xu,  Y.;  Liu,  Y.;  Zheng,  S.;  Xu,  K.;  Hu,   L.;   Wang,   C.,   Electrospun   Sb/C   Fibers   for   a   Stable   and   Fast   Sodium-­‐Ion  Battery  Anode.  ACS  Nano  2013,  7,  6378-­‐6386.   (40)   He,   M.;   Kravchyk,   K.;   Walter,   M.;   Kovalenko,   M.   V.,   Monodisperse   Antimony   Nanocrystals   for   High-­‐Rate   Li-­‐ion   and   Na-­‐ion   Battery   Anodes:   Nano   versus   Bulk.   Nano   Lett.   2014,   14,   1255-­‐1262.   (41)   Ji,  L.;  Gu,  M.;  Shao,  Y.;  Li,  X.;  Engelhard,  M.  H.;  Arey,  B.   W.;   Wang,   W.;   Nie,   Z.;   Xiao,   J.;   Wang,   C.;   Zhang,   J.-­‐G.;   Liu,   J.,   Controlling   SEI   Formation   on   SnSb-­‐Porous   Carbon   Nanofibers   for  Improved  Na  Ion  Storage.  Adv.  Mater.  2014,  26,  2901-­‐2908.   (42)   Su,   D.;   Ahn,   H.-­‐J.;   Wang,   G.,   SnO2@Graphene   Nanocomposites   as   Anode   Materials   for   Na-­‐Ion   Batteries   with   Superior   Electrochemical   Performance.   Chem.   Commun.   2013,   49,  3131-­‐3133.  

(43)   Wang,  Y.;  Su,  D.;  Wang,  C.;  Wang,  G.,  SnO2@MWCNT   Nanocomposite  as  a  High  Capacity  Anode  Material  for  Sodium-­‐ Ion  Batteries.  Electrochem.  Commun.  2013,  29,  8-­‐11.   (44)   Wu,  L.;  Hu,  X.;  Qian,  J.;  Pei,  F.;  Wu,  F.;  Mao,  R.;  Ai,  X.;   Yang,   H.;   Cao,   Y.,   A   Sn-­‐SnS-­‐C   Nanocomposite   as   Anode   Host   Materials   for   Na-­‐Ion   Batteries.   J.   Mater.   Chem.   A   2013,   1,   7181-­‐ 7184.   (45)   Qu,  B.;  Ma,  C.;  Ji,  G.;  Xu,  C.;  Xu,  J.;  Meng,  Y.  S.;  Wang,   T.;  Lee,  J.  Y.,  Layered  SnS2-­‐Reduced  Graphene  Oxide  Composite   –   A   High-­‐Capacity,   High-­‐Rate,   and   Long-­‐Cycle   Life   Sodium-­‐Ion   Battery  Anode  Material.  Adv.  Mater.  2014,  26,  3854-­‐3859.   (46)   Xie,   X.;   Su,   D.;   Chen,   S.;   Zhang,   J.;   Dou,   S.;   Wang,   G.,   SnS2  Nanoplatelet@Graphene  Nanocomposites  as  High-­‐Capacity   Anode   Materials   for   Sodium-­‐Ion   Batteries.   Chem.   Asian   J.   2014,   9,  1611-­‐1617.   (47)   Sun,   Q.;   Ren,   Q.-­‐Q.;   Li,   H.;   Fu,   Z.-­‐W.,   High   capacity   Sb2O4   Thin   Film   Electrodes   for   Rechargeable   Sodium   Battery.   Electrochem.  Commun.  2011,  13,  1462-­‐1464.   (48)   Yu,   D.   Y.   W.;   Prikhodchenko,   P.   V.;   Mason,   C.   W.;   Batabyal,  S.  K.;  Gun,  J.;  Sladkevich,  S.;  Medvedev,  A.  G.;  Lev,  O.,   High-­‐Capacity   Antimony   Sulphide   Nanoparticle-­‐Decorated   Graphene   Composite   as   Anode   for   Sodium-­‐Ion   Batteries.   Nat.   Commun.  2013,  4,  2922.   (49)   Kim,  Y.;  Park,  Y.;  Choi,  A.;  Choi,  N.-­‐S.;  Kim,  J.;  Lee,  J.;   Ryu,   J.   H.;   Oh,   S.   M.;   Lee,   K.   T.,   An   Amorphous   Red   Phosphorus/Carbon   Composite   as   a   Promising   Anode   Material   for  Sodium  Ion  Batteries.  Adv.  Mater.  2013,  25,  3045-­‐3049.   (50)   Qian,   J.;   Wu,   X.;   Cao,   Y.;   Ai,   X.;   Yang,   H.,   High   Capacity   and   Rate   Capability   of   Amorphous   Phosphorus   for   Sodium  Ion  Batteries.  Angew.  Chem.  Int.  Ed.  2013,  52,  4633-­‐4636.   (51)   Li,   W.-­‐J.;   Chou,   S.-­‐L.;   Wang,   J.-­‐Z.;   Liu,   H.-­‐K.;   Dou,   S.-­‐ X.,   Simply   Mixed   Commercial   Red   Phosphorus   and   Carbon   Nanotube   Composite   with   Exceptionally   Reversible   Sodium-­‐Ion   Storage.  Nano  Lett.  2013,  13,  5480-­‐5484.   (52)   Li,   W.;   Chou,   S.-­‐L.;   Wang,   J.-­‐Z.;   Kim,   J.   H.;   Liu,   H.-­‐K.;   Dou,   S.-­‐X.,   Sn4+xP3  @   Amorphous   Sn-­‐P   Composites   as   Anodes   for   Sodium-­‐Ion   Batteries   with   Low   Cost,   High   Capacity,   Long   Life,   and  Superior  Rate  Capability.  Adv.  Mater.  2014,  26,  4037-­‐4042.   (53)   Qian,  J.;  Xiong,  Y.;  Cao,  Y.;  Ai,  X.;  Yang,  H.,  Synergistic   Na-­‐Storage   Reactions   in   Sn4P3   as   a   High-­‐Capacity,   Cycle-­‐stable   Anode  of  Na-­‐Ion  Batteries.  Nano  Lett.  2014,  14,  1865-­‐1869.   (54)   Kim,  Y.;  Kim,  Y.;  Choi,  A.;  Woo,  S.;  Mok,  D.;  Choi,  N.-­‐ S.;  Jung,  Y.  S.;  Ryu,  J.  H.;  Oh,  S.  M.;  Lee,  K.  T.,  Tin  Phosphide  as  a   Promising   Anode   Material   for   Na-­‐Ion   Batteries.   Adv.   Mater.   2014,  26,  4139-­‐4144.   (55)   Fullenwarth,   J.;   Darwiche,   A.;   Soares,   A.;   Donnadieu,   B.;  Monconduit,  L.,  NiP3:  A  Promising  Negative  Electrode  for  Li-­‐   and  Na-­‐Ion  Batteries.  J.  Mater.  Chem.  A  2014,  2,  2050-­‐2059.   (56)   Kovnir,  K.  A.;  Kolen’ko,  Y.  V.;  Ray,  S.;  Li,  J.;  Watanabe,   T.;  Itoh,  M.;  Yoshimura,  M.;  Shevelkov,  A.  V.,  A  Facile  High-­‐Yield   Solvothermal  Route  to  Tin  Phosphide  Sn4P3.  J.  Solid  State  Chem.   2006,  179,  3756-­‐3762.   (57)   Fox,  M.,  Dangerous  When  Wet  Materials  and  Division   4.3.   In   Glossary   for   the   Worldwide   Transportation   of   Dangerous   Goods   and   Hazardous   Materials,   Springer   Berlin   Heidelberg:   1999;  pp  58-­‐60.   (58)   Longwell,   J.   P.,   Alternative   Technologies   for   the   Destruction   of   Chemical   Agents   and   Munitions.   National   Academy  Press:  1993.   (59)   Yu,   L.   H.;   Xi,   S.   B.;   Wei,   C.;   Zhang,   W.   Y.;   Du,   Y.   H.;   Yan,   Q.   Y.;   Xu,   Z.   C.,   Superior   Lithium   Storage   Properties   of   β-­‐ FeOOH.  Adv.  Energy  Mater.  2015,  5,  1401517.   (60)   Cabana,   J.;   Monconduit,   L.;   Larcher,   D.;   Rosa   Palacin,   M.,   Beyond   Intercalation-­‐Based  Li-­‐Ion  Batteries:  The  State  of  the   Art   and   Challenges   of   Electrode   Materials   Reacting   Through   Conversion  Reactions.  Adv.  Mater.  2010,  22,  E170-­‐E192.  

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(61)   Huang,   B.;   Tai,   K.;   Zhang,   M.;   Xiao,   Y.;   Dillon,   S.   J.,   Comparative  Study  of  Li  and  Na  Electrochemical  Reactions  with   Iron  Oxide  Nanowires.  Electrochim.  Acta  2014,  118,  143-­‐149.   (62)   Oh,  S.-­‐M.;  Myung,  S.-­‐T.;  Yoon,  C.  S.;  Lu,  J.;  Hassoun,  J.;   Scrosati,   B.;   Amine,   K.;   Sun,   Y.-­‐K.,   Advanced   Na[Ni0.25Fe0.5Mn0.25]O2/C–Fe3O4  Sodium-­‐Ion  Batteries  Using  EMS   Electrolyte  for  Energy  Storage.  Nano  Lett.  2014,  14,  1620-­‐1626.   (63)   Valvo,   M.;   Lindgren,   F.;   Lafont,   U.;   Bjorefors,   F.;   Edstrom,   K.,   Towards   More   Sustainable   Negative   Electrodes   in   Na-­‐Ion  Batteries  via  Nanostructured  Iron  Oxide.  J.  Power  Sources   2014,  245,  967-­‐978.   (64)   Hariharan,   S.;   Saravanan,   K.;   Ramar,   V.;   Balaya,   P.,   A   Rationally   Designed   Dual   Role   Anode   Material   for   Lithium-­‐Ion   and   Sodium-­‐Ion   Batteries:   Case   Study   of   Eco-­‐Friendly   Fe3O4.   Phys.  Chem.  Chem.  Phys.  2013,  15,  2945-­‐2953.   (65)   Jian,   Z.;   Zhao,   B.;   Liu,   P.;   Li,   F.;   Zheng,   M.;   Chen,   M.;   Shi,   Y.;   Zhou,   H.,   Fe2O3   Nanocrystals   Anchored   onto   Graphene   Nanosheets   as   the   Anode   Material   for   Low-­‐Cost   Sodium-­‐Ion   Batteries.  Chem.  Commun.  2014,  50,  1215-­‐1217.   (66)   Zhang,  N.;  Han,  X.;  Liu,  Y.;  Hu,  X.;  Zhao,  Q.;  Chen,  J.,   3D   Porous   γ-­‐Fe2O3@C   Nanocomposite   as   High-­‐Performance   Anode   Material   of   Na-­‐Ion   Batteries.   Adv.   Energy   Mater.   2014,   1401123.  

Page 10 of 11

(67)   Lopez,   M.   C.;   Lavela,   P.;   Ortiz,   G.   F.;   Tirado,   J.   L.,   Transition  Metal  Oxide  Thin  Films  With  Improved  Reversibility   as   Negative   Electrodes   for   Sodium-­‐Ion   Batteries.   Electrochem.   Commun.  2013,  27,  152-­‐155.   (68)   Jiang,  Y.;  Hu,  M.;  Zhang,  D.;  Yuan,  T.;  Sun,  W.;  Xu,  B.;   Yan,  M.,  Transition  Metal  Oxides  for  High  Performance  Sodium   Ion  Battery  Anodes.  Nano  Energy  2014,  5,  60-­‐66.   (69)   Taberna,  L.;  Mitra,  S.;  Poizot,  P.;  Simon,  P.;  Tarascon,  J.   M.,   High   Rate   Capabilities   Fe3O4-­‐Based   Cu   Nano-­‐Architectured   Electrodes   for   Lithium-­‐Ion   Battery   Applications.   Nat.   Mater.   2006,  5,  567-­‐573.   (70)   Yu,   L.   H.;   Wei,   C.;   Yan,   Q.   Y.;   Xu,   Z.   C.,   Controlled   Synthesis   of   High-­‐Performance   β-­‐FeOOH   Anodes   for   Lithium-­‐ Ion   Batteries   and   Their   Size   Effects.   Nano   Energy   2015,   13,   397-­‐ 404.   (71)   Delmer,   O.;   Balaya,   P.;   Kienle,   L.;   Maier,   J.,   Enhanced   Potential   of   Amorphous   Electrode   Materials:   Case   Study   of   RuO2.  Adv.  Mater.  2008,  20,  501-­‐505.   (72)   Conway,   B.   E.,   Transition   from   “Supercapacitor”   to   “Battery”   Behavior   in   Electrochemical   Energy   Storage.   J.   Electrochem.  Soc.  1991,  138,  1539-­‐1548.      

 

ACS Paragon Plus Environment

Page 11 of 11

3.0

1st 2nd 3rd 4th 5th

2.5 2.0 1.5 1.0 0.5 0.0 0

200 400 600 800 Capacity (mAh g-1)

Normalized absorption (a.u.)

Authors   are   required   to   submit   a   graphic   entry   for   the   Table   of   Contents   (TOC)   that,   in   conjunction   with   the   man-­‐ uscript  title,  should  give  the  reader  a  representative  idea  of  one  of  the  following:  A  key  structure,  reaction,  equation,   concept,   or   theorem,   etc.,   that   is   discussed   in   the   manuscript.   Consult   the   journal’s   Instructions   for   Authors   for   TOC  graphic  specifications.  

Potential (V, vs. Na +/Na)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

1st discharge

(0) Fe (III) 7120

β-FeOOH 1.1 V

0.8 V 0.3 V 0.005 V Fe

7140 7160 Energy (eV)

 

Insert  Table  of  Contents  artwork  here    

ACS Paragon Plus Environment

11